4 The Sperner property.

advertisement
4
The Sperner property.
In this section we consider a surprising application of certain adjacency ma­
trices to some problems in extremal set theory. An important role will also
be played by finite groups. In general, extremal set theory is concerned with
finding (or estimating) the most or least number of sets satisfying given settheoretic or combinatorial conditions. For example, a typical easy problem
in extremal set theory is the following: What is the most number of subsets
of an n-element set with the property that any two of them intersect? (Can
you solve this problem?) The problems to be considered here are most con­
veniently formulated in terms of partially ordered sets, or posets for short.
Thus we begin with discussing some basic notions concerning posets.
4.1 Definition. A poset (short for partially ordered set) P is a finite
set, also denoted P , together with a binary relation denoted � satisfying the
following axioms:
(P1) (reflexivity) x � x for all x √ P
(P2) (antisymmetry) If x � y and y � x, then x = y.
(P3) (transitivity) If x � y and y � z, then x � z.
One easy way to obtain a poset is the following. Let P be any collection
of sets. If x, y √ P , then define x � y in P if x ∪ y as sets. It is easy to see
that this definition of � makes P into a poset. If P consists of all subsets
of an n-element set S, then P is called a (finite) boolean algebra of rank n
and is denoted by BS . If S = {1, 2, . . . , n}, then we denote BS simply by Bn .
Boolean algebras will play an important role throughout this section.
There is a simple way to represent small posets pictorially. The Hasse
diagram of a poset P is a planar drawing, with elements of P drawn as dots.
If x < y in P (i.e., x � y and x ∈= y), then y is drawn “above” x (i.e.,
with a larger vertical coordinate). An edge is drawn between x and y if
y covers x, i.e., x < y and no element z is in between, i.e., no z satisfies
x < z < y. By the transitivity property (P3), all the relations of a finite
17
poset are determined by the cover relations, so the Hasse diagram determines
P . (This is not true for infinite posets; for instance, the real numbers R with
their usual order is a poset with no cover relations.) The Hasse diagram of
the boolean algebra B3 looks like
� 123
��
��
�
�
�
� 12 � 13 �� 23
� ��
�� �
�
�
�
� 1 ���2 �� 3
�
�
�
�
���
Ø
We say that two posets P and Q are isomorphic if there is a bijection
(one-to-one and onto function) � : P � Q such that x � y in P if and only
if �(x) � �(y) in Q. Thus one can think that two posets are isomorphic if
they differ only in the names of their elements. This is exactly analogous to
the notion of isomorphism of groups, rings, etc. It is an instructive exercise
to draw Hasse diagrams of the one poset of order (number of elements) one
(up to isomorphism), the two posets of order two, the five posets of order
three, and the sixteen posets of order four. More ambitious readers can try
the 63 posets of order five, the 318 of order six, the 2045 of order seven, the
16999 of order eight, the 183231 of order nine, the 2567284 of order ten, the
46749427 of order eleven, the 1104891746 of order twelve, the 33823827452
of order thirteen, and the 1338193159771 of order fourteen. Beyond this the
number is not currently known.
A chain C in a poset is a totally ordered subset of P , i.e., if x, y √ C then
either x � y or y � x in P . A finite chain is said to have length n if it has
n + 1 elements. Such a chain thus has the form x0 < x1 < · · · < xn . We say
that a finite poset is graded of rank n if every maximal chain has length n.
(A chain is maximal if it’s contained in no larger chain.) For instance, the
boolean algebra Bn is graded of rank n [why?]. A chain y0 < y1 < · · · < yj is
said to be saturated if each yi+1 covers yi . Such a chain need not be maximal
since there can be elements of P smaller than y0 or greater than yj . If P is
graded of rank n and x √ P , then we say that x has rank j, denoted α(x) = j,
if some (or equivalently, every) saturated chain of P with top element x has
length j. Thus [why?] if we let Pj = {x √ P : α(x) = j}, then P is a
18
disjoint union P = P0 ≤ P1 ≤ · · · ≤ Pn , and every maximal chain of P has the
form x0 < x1 < · · · < xn where α(xj ) = j. We write pj = |Pj |, the number
of elements of P of rank j. For example, if P = Bn then α(x) = |x| (the
cardinality of x as a set) and
� �
n
pj = #{x ∪ {1, 2, · · · , n} : |x| = j} =
.
j
(Note that we use both |S | and #x for the cardinality of the finite set S.)
We say that a graded poset P of rank n (always assumed to be finite)
is rank-symmetric if pi = pn−i for 0 � i � n, and rank-unimodal if p0 �
p1 � · · · � pj ← pj+1 ← pj+2 ← · · · ← pn for some 0 � j � n. If P is both
rank-symmetric and rank-unimodal, then we clearly have
p0 � p1 � · · · � pm ← pm+1 ← · · · ← pn , if n = 2m
p0 � p1 � · · · � p
m = pm+1 ← pm+2 ← · · · ← pn , if n = 2m + 1.
We also say that the sequence p0 , p1 , . . . , pn itself or the polynomial F (q) =
p0 + p1 q + · · · + pn q n is symmetric or unimodal, as the case may be. For
instance, Bn is rank-symmetric and rank-unimodal,
�n� �n�
�n�since it is well-known
(and easy to prove) that the sequence 0 , 1 , . . . , n (the nth row of Pas­
cal’s triangle) is symmetric and unimodal. Thus the polynomial (1 + q)n is
symmetric and unimodal.
A few more definitions, and then finally some results! An antichain in
a poset P is a subset A of P for which no two elements are comparable,
i.e., we can never have x, y √ A and x < y. For instance, in a graded
poset P the “levels” Pj are antichains [why?]. We will be concerned with the
problem of finding the largest antichain in a poset. Consider for instance the
boolean algebra Bn . The problem of finding the largest antichain in Bn is
clearly equivalent to the following problem in extremal set theory: Find the
largest collection of subsets of an n-element set such that no element of the
collection contains another. A good guess would be to take all the subsets
of cardinality
� n �⊂n/2⊆ (where ⊂x⊆ denotes the greatest integer � x), giving a
total of ≥n/2∈
sets in all. But how can we actually prove there is no larger
collection? Such a proof was first given by Emmanuel Sperner in 1927 and
is known as Sperner’s theorem. We will give two proofs of Sperner’s theorem
19
in this section; one proof uses linear algebra and will be applied to certain
other situations, while the other proof is an elegant combinatorial argument
due to David Lubell in 1966, which we present for its “cultural value.” Our
extension of Sperner’s theorem to certain other situations will involve the
following crucial definition.
4.2 Definition. Let P be a graded poset of rank n. We say that P
has the Sperner property or is a Sperner poset if
max{|A| : A is an antichain of P } = max{|Pi | : 0 � i � n}.
In other words, no antichain is larger than the largest level Pi .
Thus Sperner’s theorem is equivalent to saying that Bn has the Sperner
property. Note that if P has the Sperner property there may still be an­
tichains of maximum cardinality other than the biggest Pi ; there just can’t
be any bigger antichains.
4.3 Example. A simple example of a graded poset that fails to satisfy
the Sperner property is the following:
�
��
����
�
�� �
��
�
�
�
�
We now will discuss a simple combinatorial condition which guarantees
that certain graded posets P are Sperner. We define an order-matching from
Pi to Pi+1 to be a one-to-one function µ : Pi � Pi+1 satisfying x < µ(x)
for all x √ Pi . Clearly if such an order-matching exists then pi � pi+1
(since µ is one-to-one). Easy examples show that the converse is false, i.e.,
if pi � pi+1 then there need not exist an order-matching from Pi to Pi+1 .
We similarly define an order-matching from Pi to Pi−1 to be a one-to-one
function µ : Pi � Pi−1 satisfying µ(x) < x for all x √ Pi .
4.4 Proposition.
Let P be a graded poset of rank n. Suppose there
exists an integer 0 � j � n and order-matchings
P0 � P1 � P2 � · · · � Pj � Pj+1 � Pj+2 � · · · � Pn .
Then P is rank-unimodal and Sperner.
20
(17)
Proof. Since order-matchings are one-to-one it is clear that
p0 � p1 � · · · � pj ← pj+1 ← pj+2 ← · · · ← pn .
Hence P is rank-unimodal.
Define a graph G as follows. The vertices of G are the elements of P . Two
vertices x, y are connected by an edge if one of the order-matchings µ in the
statement of the proposition satisfies µ(x) = y. (Thus G is a subgraph of the
Hasse diagram of P .) Drawing a picture will convince you that G consists of
a disjoint union of paths, including single-vertex paths not involved in any
of the order-matchings. The vertices of each of these paths form a chain in
P . Thus we have partitioned the elements of P into disjoint chains. Since P
is rank-unimodal with biggest level Pj , all of these chains must pass through
Pj [why?]. Thus the number of chains is exactly pj . Any antichain A can
intersect each of these chains at most once, so the cardinality |A| of A cannot
exceed the number of chains, i.e., |A| � pj . Hence by definition P is Sperner.
�
It is now finally time to bring some linear algebra into the picture. For
any (finite) set S, we let RS denote the real vector space consisting of all
formal linear combinations (with real coefficients) of elements of S. Thus S
is a basis for RS, and in fact we could have simply defined RS to be the
real vector space with basis S. The next lemma relates the combinatorics we
have just discussed to linear algebra and will allow us to prove that certain
posets are Sperner by the use of linear algebra (combined with some finite
group theory).
4.5 Lemma. Suppose there exists a linear transformation U : RPi �
RPi+1 (U stands for “up”) satisfying:
• U is one-to-one.
• For all x √ Pi , U (x) is a linear combination of elements y √ Pi+1
satisfying x < y. (We then call U an order-raising operator.)
Then there exists an order-matching µ : Pi � Pi+1 .
21
Similarly, suppose there exists a linear transformation U : RPi � RPi+1
satisfying:
• U is onto.
• U is an order-raising operator.
Then there exists an order-matching µ : Pi+1 � Pi .
Proof. Suppose U : RPi � RPi+1 is a one-to-one order-raising operator.
Let [U ] denote the matrix of U with respect to the bases Pi of RPi and Pi+1
of RPi+1 . Thus the columns of [U ] are indexed by the elements x1 , . . . , xpi of
Pi (in some order) and the rows by the elements y1 , . . . , ypi+1 of Pi+1 . Since
U is one-to-one, the rank of [U ] is equal to pi (the number of columns). Since
the column rank of a matrix equals its row rank, [U ] must have pi linearly
independent rows. Say we have labelled the elements of Pi+1 so that the first
pi rows of [U ] are linearly independent.
Let A = (aij ) be the pi × pi matrix whose rows are the first pi rows of
[U ]. (Thus A is a square submatrix of [U ].) Since the rows of A are linearly
independent, we have
�
det(A) =
±a�(1),1 · · · a�(pi ),pi ∈= 0,
where the sum is over all permutations λ of 1, . . . , pi . Thus some term
±a�(1),1 · · · a�(pi ),pi of the above sum in nonzero. Since U is order-raising, this
means that [why?] xk < y�(k) for 1 � k � pi . Hence the map µ : Pi � Pi+1
defined by µ(xk ) = y�(k) is an order-matching, as desired.
The case when U is onto rather than one-to-one is proved by a completely
analogous argument. �
We now want to apply Proposition 4.4 and Lemma 4.5 to the boolean
algebra Bn . For each 0 � i < n, we need to define a linear transformation
Ui : R(Bn )i � R(Bn )i+1 , and then prove it has the desired properties. We
simply define Ui to be the simplest possible order-raising operator, namely,
22
for x √ (Bn )i , let
Ui (x) =
�
y.
(18)
y∗(Bn )i+1
y>x
Note that since (Bn )i is a basis for R(Bn )i , equation (18) does indeed define
a unique linear transformation Ui : R(Bn )i � R(Bn )i+1 . By definition Ui is
order-raising; we want to show that Ui is one-to-one for i < n/2 and onto for
i ← n/2. There are several ways to show this using only elementary linear
algebra; we will give what is perhaps the simplest proof, though it is quite
tricky. The idea is to introduce “dual” operators Di : R(Bn )i � (Bn )i−1 to
the Ui ’s (D stands for “down”), defined by
�
Di (y) =
x,
(19)
x∗(Bn )i−1
x<y
for all y √ (Bn )i . Let [Ui ] denote the matrix of Ui with respect to the bases
(Bn )i and (Bn )i+1 , and similarly let [Di ] denote the matrix of Di with respect
to the bases (Bn )i and (Bn )i−1 . A key observation which we will use later is
that
[Di+1 ] = [Ui ]t ,
(20)
i.e., the matrix [Di+1 ] is the transpose of the matrix [Ui ] [why?]. Now let
Ii : R(Bn )i � R(Bn )i denote the identity transformation on R(Bn )i , i.e.,
Ii (u) = u for all u √ R(Bn )i . The next lemma states (in linear algebraic
terms) the fundamental combinatorial property of Bn which we need. For
this lemma set Un = 0 and D0 = 0 (the 0 linear transformation between the
appropriate vector spaces).
4.6 Lemma.
Let 0 � i � n. Then
Di+1 Ui − Ui−1 Di = (n − 2i)Ii .
(21)
(Linear transformations are multiplied right-to-left, so AB(u) = A(B(u)).)
Proof. Let x √ (Bn )i . We need to show that if we apply the left-hand
side of (21) to x, then we obtain (n − 2i)x. We have
⎞
�
� � ⎜
y⎝
Di+1 Ui (x) = Di+1 �
|y|=i+1
x�y
23
=
� �
z.
|y|=i+1 |z|=i
x�y
z�y
If x, z √ (Bn )i satisfy |x → z | < i − 1, then there is no y √ (Bn )i+1 such that
x ∩ y and z ∩ y. Hence the coefficient of z in Di+1 Ui (x) when it is expanded
in terms of the basis (Bn )i is 0. If |x → z | = i − 1, then there is one such y,
namely, y = x ≤ z. Finally if x = z then y can be any element of (Bn )i+1
containing x, and there are n − i such y in all. It follows that
�
z.
(22)
Di+1 Ui (x) = (n − i)x +
|z|=i
|x�z|=i−1
By exactly analogous reasoning (which the reader should check), we have for
x √ (Bn )i that
�
z.
(23)
Ui−1 Di (x) = ix +
|z|=i
|x�z|=i−1
Subtracting (23) from (22) yields (Di+1 Ui −Ui−1 Di )(x) = (n−2i)x, as desired.
�
4.7 Theorem. The operator Ui defined above is one-to-one if i < n/2
and is onto if i ← n/2.
Proof. Recall that [Di ] = [Ui−1 ]t . From linear algebra we know that
a (rectangular) matrix times its transpose is positive semidefinite (or just
semidefinite for short) and hence has nonnegative (real) eigenvalues. By
Lemma 4.6 we have
Di+1 Ui = Ui−1 Di + (n − 2i)Ii .
Thus the eigenvalues of Di+1 Ui are obtained from the eigenvalues of Ui−1 Di
by adding n − 2i. Since we are assuming that n − 2i > 0, it follows that the
eigenvalues of Di+1 Ui are strictly positive. Hence Di+1 Ui is invertible (since
it has no 0 eigenvalues). But this implies that Ui is one-to-one [why?], as
desired.
The case i ← n/2 is done by a “dual” argument (or in fact can be deduced
directly from the i < n/2 case by using the fact that the poset Bn is “self­
dual,” though we will not go into this). Namely, from the fact that
Ui Di+1 = Di+2 Ui+1 + (2i + 2 − n)Ii+1
24
we get that Ui Di+1 is invertible, so now Ui is onto, completing the proof. �
Combining Proposition 4.4, Lemma 4.5, and Theorem 4.7, we obtain the
famous theorem of Sperner.
4.8 Corollary.
The boolean algebra Bn has the Sperner property.
It is natural to ask whether there is a less indirect proof of Corollary
4.8. In fact, several nice proofs are known; we give one due to David Lubell,
mentioned before Definition 4.2.
Lubell’s proof of Sperner’s theorem. First we count the total number
of maximal chains Ø = x0 < x1 < · · · < xn = {1, . . . , n} in Bn . There are n
choices for x1 , then n − 1 choices for x2 , etc., so there are n! maximal chains
in all. Next we count the number of maximal chains x0 < x1 < · · · < xi =
x < · · · < xn which contain a given element x of rank i. There are i choices
for x1 , then i − 1 choices for x2 , up to one choice for xi . Similarly there are
n − i choices for xi+1 , then n − 2 choices for xi+2 , etc., up to one choice for
xn . Hence the number of maximal chains containing x is i!(n − i)!.
Now let A be an antichain. If x √ A, then let Cx be the set of maximal
chains of Bn which contain x. Since A is an antichain, the sets Cx , x √ A
are pairwise disjoint. Hence
�
�
|
Cx | =
|Cx |
x�A
x�A
=
�
x�A
(α(x))!(n − α(x))!
Since the total number of maximal chains in the Cx ’s cannot exceed the total
number n! of maximal chains in Bn , we have
�
(α(x))!(n − α(x))! � n!
x�A
Divide both sides by n! to obtain
�
x�A
1
n
λ(x)
�
� � 1.
25
Since
� n�
i
is maximized when i = ⊂n/2⊆, we have
1
�
n
≥n/2∈
for all x √ A (or all x √ Bn ). Thus
�
x�A
or equivalently,
���
1
�
n
≥n/2∈
|A| �
Since
�
�
n
≥n/2∈
�
�
1
n
λ(x)
�,
� � 1,
�
n
.
⊂n/2⊆
is the size of the largest level of Bn , it follows that Bn is Sperner.
In view of the above elegant proof of Lubell, the reader may be wondering
what was the point of giving a rather complicated and indirect proof using
linear algebra. Admittedly, if all we could obtain from the linear algebra
machinery we have developed was just another proof of Sperner’s theorem,
then it would have been hardly worth the effort. But in the next section we
will show how Theorem 4.7, when combined with a little finite group theory,
can be used to obtain many interesting combinatorial results for which simple,
direct proofs are not known.
26
Download